Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Review Article
  • Published:

Polyglutamine spinocerebellar ataxias — from genes to potential treatments

Key Points

  • Among the diverse group of dominantly inherited spinocerebellar ataxias (SCAs), those attributable to the expansion of polyglutamine (polyQ)-encoding CAG repeats include the most prevalent and severe forms (SCA1, SCA2, SCA3, SCA6, SCA7 and SCA17).

  • The polyQ SCAs typically present with gait ataxia, limb incoordination, speech disturbance and oculomotor abnormalities, and death is caused by brainstem failure; however, SCA7 is uniquely characterized by retinal degeneration, and SCA6 is usually a pure cerebellar disease that does not reduce lifespan.

  • In the polyQ SCAs, the Purkinje cells of the cerebellar cortex are a prominent pathological target. An exception is SCA3, in which Purkinje cells are less involved.

  • Changes in the expression of receptors and ion channels important for regulating membrane excitability contribute to motor dysfunction, as well as to structural changes in neurons that lead to cell death and thus may be targets for the treatment of motor dysfunction.

  • The diverse biological functions of the polyQ SCA proteins, which include regulation of transcription, RNA splicing and metabolism, and deubiquitinase activity, help to specify the disease pathogenesis of each disease.

  • Several cellular pathways are implicated in the pathogenesis of each polyQ SCA; hence, developing therapies that directly target the expression of the mutant gene or protein is a major current focus of research in this field.

Abstract

The dominantly inherited spinocerebellar ataxias (SCAs) are a large and diverse group of neurodegenerative diseases. The most prevalent SCAs (SCA1, SCA2, SCA3, SCA6 and SCA7) are caused by expansion of a glutamine-encoding CAG repeat in the affected gene. These SCAs represent a substantial portion of the polyglutamine neurodegenerative disorders and provide insight into this class of diseases as a whole. Recent years have seen considerable progress in deciphering the clinical, pathological, physiological and molecular aspects of the polyglutamine SCAs, with these advances establishing a solid base from which to pursue potential therapeutic approaches.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Figure 1: Components of the cerebellar circuitry showing degeneration in the polyglutamine spinocerebellar ataxias.
Figure 2: Cellular processes affected by mutant polyglutamine proteins in spinocerebellar ataxias.
Figure 3: Functional motifs in ATXN1, ATXN2 and ATXN3.
Figure 4: Alterations in Purkinje cell electrophysiology in spinocerebellar ataxias.

Similar content being viewed by others

References

  1. Durr, A. Autosomal dominant cerebellar ataxias: polyglutamine expansions and beyond. Lancet Neurol. 9, 885–894 (2010).

    Article  CAS  PubMed  Google Scholar 

  2. Monin, M.-L. et al. Survival and severity in dominant cerebellar ataxias. Ann. Clin. Transl. Neurol. 2, 202–207 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  3. Klockgether, T. et al. The natural history of degenerative ataxia: a retrospective study in 466 patients. Brain 121, 589–600 (1998).

    Article  PubMed  Google Scholar 

  4. Globas, C. et al. Early symptoms in spinocerebellar ataxia type 1, 2, 3, and 6. Mov. Disord. 23, 2232–2238 (2008).

    Article  PubMed  Google Scholar 

  5. Luo, L. et al. The initial symptom and motor progression in spinocerebellar ataxias. Cerebellum 16, 616–622 (2010).

    Google Scholar 

  6. Schols, L., Linnemann, C. & Globas, C. Electrophysiology in spinocerebellar ataxias: spread of disease and characteristic findings. Cerebellum 7, 198–203 (2008).

    Article  PubMed  CAS  Google Scholar 

  7. Liang, L., Chen, T. & Wu, Y. The electrophysiology of spinocerebellar ataxias. Neurophysiol. Clin. 46, 27–34 (2016).

    Article  PubMed  Google Scholar 

  8. Paulson, H. Machado–Joseph disease/spinocerebellar ataxia type 3. Handb. Clin. Neurol. 103, 437–449 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  9. Seidel, K. et al. Brain pathology of spinocerebellar ataxias. Acta Neuropathol. 124, 1–21 (2012).

    Article  CAS  PubMed  Google Scholar 

  10. Riess, O. et al. SCA3: neurological features, pathogenesis and animal models. Cerebellum 7, 125–137 (2008).

    Article  CAS  PubMed  Google Scholar 

  11. Jacobi, H. et al. The natural history of spinocerebellar ataxia type 1, 2, 3, and 6: a 2-year follow-up study. Neurology 77, 1035–1041 (2011). One of several recent natural history studies performed by the EUROSCA that will inform future clinical prevention trials in polyQ SCAs (also see Ref. 130).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Ashizawa, T. et al. Clinical characteristics of patients with spinocerebellar ataxias 1, 2, 3 and 6 in the US; a prospective observational study. Orphanet. J. Rare Dis. 8, 177 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  13. Robitaille, Y., Schut, L. & Kish, S. J. Structural and immunocytochemical features of olivopontocerebellar atrophy caused by the spinocerebellar ataxia type 1 (SCA-1) mutation define a unique phenotype. Acta Neuropathol. 90, 572–581 (1995).

    Article  CAS  PubMed  Google Scholar 

  14. Orozco, G. et al. Dominantly inherited olivopontocerebellar atrophy from eastern Cuba. Clinical, neuropathological, and biochemical findings. J. Neurol. Sci. 93, 37–50 (1989).

    Article  CAS  PubMed  Google Scholar 

  15. Estrada, R. et al. Spinocerebellar ataxia 2 (SCA2): morphometric analyses in 11 autopsies. Acta Neuropathol. 97, 306–310 (1999).

    Article  CAS  PubMed  Google Scholar 

  16. Adams, C., Starkman, S. & Pulst, S. M. Clinical and molecular analysis of a pedigree of southern Italian ancestry with spinocerebellar ataxia type 2. Neurology 49, 1163–1166 (1997).

    Article  CAS  PubMed  Google Scholar 

  17. Elden, A. C. et al. Ataxin-2 intermediate-length polyglutamine expansions are associated with increased risk for ALS. Nature 466, 1069–1075 (2010). This study demonstrates that intermediate-length alleles are a risk factor for ALS, indicating that even non-pathogenic changes in repeat length can have profound effects on neuronal function.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Van Damme, P. et al. Expanded ATXN2 CAG repeat size in ALS identifies genetic overlap between ALS and SCA2. Neurology 76, 2066–2072 (2011).

    Article  CAS  PubMed  Google Scholar 

  19. Tan, R. H. et al. Cerebellar neuronal loss in amyotrophic lateral sclerosis cases with ATXN2 intermediate repeat expansions. Ann. Neurol. 79, 295–305 (2016).

    Article  CAS  PubMed  Google Scholar 

  20. Sesqueiros, J. & Coutinho, P. Epidemiology and clinical aspects of Machado–Joseph disease. Adv. Neurol. 61, 139–153 (1993).

    Google Scholar 

  21. Rüb, U., Brunt, E. R. & Deller, T. New insights into the pathoanatomy of spinocerebellar ataxia type 3 (Machado–Joseph disease). Curr. Opin. Neurol. 21, 111–116 (2008).

    Article  PubMed  Google Scholar 

  22. Rüb, U. et al. Clinical features, neurogenetics and neuropathology of the polyglutamine spinocerebellar ataxias type 1, 2, 3, 6 and 7. Prog. Neurobiol. 104, 38–66 (2013).

    Article  PubMed  CAS  Google Scholar 

  23. Stevanin, G. et al. Clinical and molecular features of spinocerebellar ataxia type 6. Neurology 49, 1243–1246 (1997).

    Article  CAS  PubMed  Google Scholar 

  24. Schulz, J. B. et al. Visualization, quantification and correlation of brain atrophy with clinical symptoms in spinocerebellar ataxia types 1, 3 and 6. Neuroimage 49, 158–168 (2010).

    Article  PubMed  Google Scholar 

  25. Gierga, K. et al. Spinocerebellar ataxia type 6 (SCA6): neurodegeneration goes beyond the known brain predilection sites. Neuropathol. Appl. Neurobiol. 35, 515–527 (2009).

    Article  CAS  PubMed  Google Scholar 

  26. Shao, J. & Diamond, M. I. Polyglutamine diseases: emerging concepts in pathogenesis and therapy. Hum. Mol. Genet. 16, R115–R123 (2007).

    Article  CAS  PubMed  Google Scholar 

  27. Klement, I. A. et al. Ataxin-1 nuclear localization and aggregation: role in polyglutamine-induced disease in SCA1 transgenic mice. Cell 95, 41–53 (1998).

    Article  CAS  PubMed  Google Scholar 

  28. Irwin, S. et al. RNA association and nucleocytoplasmic shuttling by ataxin-1. J. Cell Sci. 118, 233–242 (2005).

    Article  CAS  PubMed  Google Scholar 

  29. Yue, S. et al. The spinocerebellar ataxia type 1 protein, ataxin-1, has RNA-binding activity that is inversely affected by the length of its polyglutamine tract. Hum. Mol. Genet. 10, 25–30 (2001).

    Article  CAS  PubMed  Google Scholar 

  30. Tsai, C.-C. et al. Ataxin-1, a SCA1 neurodegenerative disorder protein, is functionally linked to the silencing mediator of retinoid and thyroid hormone receptors. Proc. Natl Acad. Sci. USA 101, 4047–4052 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Lam, Y. C. et al. ATAXIN-1 interacts with the repressor Capicua in its native complex to cause SCA1 neuropathology. Cell 127, 1335–1347 (2006).

    Article  CAS  PubMed  Google Scholar 

  32. Tsuda, H. et al. The AXH domain in mammalian/Drosophila Ataxin-1 mediates neurodegeneration in spinocerebellar ataxia 1 through its interaction with Gfi-1/Senseless proteins. Cell 122, 633–644 (2005).

    Article  CAS  PubMed  Google Scholar 

  33. Serra, H. G. et al. RORα-mediated Purkinje cell development determines disease severity in adult SCA1 mice. Cell 127, 697–708 (2006).

    Article  CAS  PubMed  Google Scholar 

  34. Gehrking, K. M. et al. Partial loss of Tip60 slows midstage neurodegeneration in a spinocerebellar ataxia type 1 (SCA1) mouse model. Hum. Mol. Genet. 20, 2204–2212 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. de Chiara, C. et al. The AXH module: an independently folded domain common to ataxin-1 and HBP1. FEBS Lett. 551, 107–112 (2003).

    Article  CAS  PubMed  Google Scholar 

  36. Chen, Y. W. et al. The structure of the AXH domain of spinocerebellar ataxin-1. J. Biol. Chem. 279, 3758–3765 (2004).

    Article  CAS  PubMed  Google Scholar 

  37. Fryer, J. D. et al. Exercise and genetic rescue of SCA1 via the transcriptional repressor Capicua. Science 334, 690–693 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Emamian, E. S. et al. Serine 776 of ataxin-1 is critical for polyglutamine-induced disease in SCA1 transgenic mice. Neuron 38, 375–387 (2003).

    Article  CAS  PubMed  Google Scholar 

  39. Huttlin, E. L. et al. A tissue-specific atlas of mouse protein phosphorylation and expression. Cell 143, 1174–1189 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Duvick, L. et al. SCA1-like disease in mice expressing wild type ataxin-1 with a serine to aspartic acid replacement at residue 776. Neuron 67, 929–935 (2010). The disease-like phenotypes elicited by the engineered ATXN1 in this study, despite a normal repeat length, indicate that specific altered protein interactions play a key part in SCA1.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Jorgensen, N. D. et al. Phosphorylation of ATXN1 at Ser776 in the cerebellum. J. Neurochem. 110, 675–686 (2009) (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Chen, H.-K. et al. Interaction of Akt-phosphorylated ataxin-1 with 14-3-3 mediates neurodegeneration in spinocerebellar ataxia type 1. Cell 113, 457–468 (2003).

    Article  CAS  PubMed  Google Scholar 

  43. Lim, J. et al. Opposing effects of polyglutamine expansion on native protein complexes contribute to SCA1. Nature 452, 713–719 (2008). The study presents data supporting the concept that an altered balance in the interaction of expanded ATXN1 with CIC and RBM17 drives pathogenesis in SCA1.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. de Chiara, C. et al. Phosphorylation of S776 and 14-3-3 binding modulate Ataxin-1 interaction with splicing factors. PLoS ONE 4, e8372 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  45. Bentley, D. L. Coupling mRNA processing with transcription in time and space. Nat. Rev. Genet. 15, 163–175 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Ingram, M. et al. Cerebellar transcriptome profiles of ATXN1 transgenic mice reveal SCA1 disease progression and protection pathways. Neuron 89, 1194–1207 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Serra, H. G. et al. Gene profiling links SCA1 pathophysiology to glutamate signaling in Purkinje cells of transgenic mice. Hum. Mol. Genet. 13, 2535–2543 (2004). One of several reports suggesting that altered glutamate signalling in the cerebellum contributes to disease in polyQ SCAs.

    Article  CAS  PubMed  Google Scholar 

  48. Carlson, K. M., Andresen, M. J. & Orr, H. T. Emerging pathogenic pathways in the spinocerebellar ataxias. Curr. Opin. Genet. Dev. 19, 247–253 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Schorge, S. et al. Human ataxias: a genetic dissection of inositol triphosphate receptor (ITPR1)-dependent signaling. Trends Neurosci. 33, 209–211 (2010).

    Article  CAS  Google Scholar 

  50. Ruegsegger, C. et al. Impaired mTORC1-dependent expression of Homer-3 influences SCA1 pathophysiology. Neuron 89, 129–146 (2016).

    Article  CAS  PubMed  Google Scholar 

  51. Sánchez, I., Balagué, E. & Matilla-Dueñas, A. Ataxin-1 regulates the cerebellar bioenergetics proteome through the GSK3β–mTOR pathway which is altered in spinocerebellar ataxia type 1 (SCA1). Hum. Mol. Genet. 25, 4021–4040 (2016).

    Article  PubMed  CAS  Google Scholar 

  52. Lee, J. H. et al. Reinstating aberrant mTORC1 activity in Huntington's disease mice improves disease phenotypes. Neuron 85, 303–315 (2015).

    Article  CAS  PubMed  Google Scholar 

  53. Ross, C. A. & Truant, R. A unifying mechanism in neurodegeneration. Nature 541, 34–35 (2017).

    Article  CAS  PubMed  Google Scholar 

  54. Bettencourt, C. et al. DNA repair pathways underlie a common genetic mechanism modulating onset in polyglutamine diseases. Ann. Neurol. 79, 983–990 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Taniguchi, J. B. et al. RpA1 ameliorates symptoms of mutant ataxin-1 knock-in mice and enhances DNA damage repair. Hum. Mol. Genet. 25, 4432–4447 (2016).

    CAS  PubMed  Google Scholar 

  56. Ito, H. et al. HMGB1 facilitates repair of mitochondrial DNA damage and extends the lifespan of mutant ataxin-1 knock-in mice. EMBO Mol. Med. 7, 78–101 (2015).

    Article  CAS  PubMed  Google Scholar 

  57. Nechiporuk, T. et al. The mouse SCA2 gene: cDNA sequence, alternative splicing and protein expression. Hum. Mol. Genet. 7, 1301–1309 (1998).

    Article  CAS  PubMed  Google Scholar 

  58. Neuwald, A. F. & Koonin, E. V. Ataxin-2, global regulators of bacterial gene expression and spliceosomal snRNP proteins share a conserved domain. J. Mol. Med. 76, 3–5 (1998).

    Article  CAS  PubMed  Google Scholar 

  59. Kozlov, G. et al. Structure and function of the C-terminal PABC domain of human poly(A)-binding protein. Proc. Natl Acad. Sci. USA 98, 4409–4413 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Shibata, H., Huynh, D. P. & Pulst, S. M. A novel protein with RNA-binding motifs interacts with ataxin-2. Hum. Mol. Genet. 9, 1303–1313 (2000).

    Article  CAS  PubMed  Google Scholar 

  61. Nonhoff, U. et al. Ataxin-2 interacts with the DEAD/H-box RNA helicase DDX6 and interferes with P-bodies and stress granules. Mol. Biol. Cell 18, 1385–1396 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. McCann, C. et al. The Ataxin-2 protein is required for microRNA function and synapse-specific long-term olfactory habituation. Proc. Natl Acad. Sci. USA 108, E655–E662 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  63. Mangue, A., Amrani, N. & Jacobson, A. Pbp1p, a factor interacting with Saccharomyces cerevisiae poly(A)-binding protein, regulates polyadenylation. Mol. Cell. Biol. 18, 7383–7396 (1998).

    Article  Google Scholar 

  64. Satterfield, T. F. & Pallanck, L. J. Ataxin-2 and its Drosophila homolog, ATX2, physically assemble with polyribosomes. Hum. Mol. Genet. 15, 2523–2532 (2006).

    Article  CAS  PubMed  Google Scholar 

  65. Takahara, T. & Maeda, T. Transient sequestration of TORC1 into stress granules during heat stress. Mol. Cell 47, 242–252 (2012).

    Article  CAS  PubMed  Google Scholar 

  66. Bar, D.Z. et al. Cell size and fat content of dietary-restricted Caenorhabditis elegans are regulated by ATX-2, an mTOR repressor. Proc. Natl Acad. Sci. USA 113, E4620–E4629 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Lastres-Becker, I. et al. Mammalian ataxin-2 modulates translation control at the pre-initiation complex via PI3K/mTOR and is induced by starvation. Biochim. Biophys. Acta 1862, 1558–1569 (2016). One of several recent studies showing that ATXN2 has a complex role in regulating translation in cells.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Lagier-Tourenne, C. & Cleveland, D. C. Rethinking ALS: the FUS about TDP-43. Cell 136, 1001–1004 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Damrath, E., et al. ATXN2–CAG42 sequesters PABPC1 into insolubility and induces FBXW8 in cerebellum of old ataxic knock-in mice. PLoS Genet. 8, e1002920 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Pflieger, L. T. et al. Gene co-expression network analysis for identifying modules and functionally enriched pathways in SCA2. Hum. Mol. Genet. http://dx.doi.org/10.1093/hmg/ddx191 (2017).

  71. Costa Mdo, C. & Paulson, H. L. Toward understanding Machado–Joseph disease. Prog. Neurobiol. 97, 239–257 (2012). A thorough review describing the molecular features of SCA3, also known as Machado–Joseph disease.

    Article  PubMed  CAS  Google Scholar 

  72. Li, X. et al. Toward therapeutic targets for SCA3: Insight into the role of Machado–Joseph disease protein ataxin-3 in misfolded proteins clearance. Prog. Neurobiol. 132, 34–58 (2015).

    Article  CAS  PubMed  Google Scholar 

  73. Faggiano, S. et al. Allosteric regulation of deubiquitylase activity through ubiquitination. Front. Mol. Biosci. 2, 2 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  74. Blount, J. R. et al. Ubiquitin-binding site 2 of ataxin-3 prevents its proteasomal degradation by interacting with Rad23. Nat. Commun. 5, 4638 (2014).

    Article  CAS  PubMed  Google Scholar 

  75. Tsou, W. L. et al. Ubiquitination regulates the neuroprotective function of the deubiquitinase ataxin-3 in vivo. J. Biol. Chem. 288, 34460–34469 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Durcan, T. M. & Fon, E. A. Ataxin-3 and its E3 partners: implications for Machado–Joseph disease. Front. Neurol. 4, 46 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  77. Scaglione, K. M. et al. Ube2w and ataxin-3 coordinately regulate the ubiquitin ligase CHIP. Mol. Cell 43, 599–612 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Wang, Q., Li, L. & Ye, Y. Regulation of retrotranslocation by p97-associated deubiquitinating enzyme ataxin-3. J. Cell. Biol. 174, 963–971 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Wang, H., Ying, Z. & Wang, G. Ataxin-3 regulates aggresome formation of copper-zinc superoxide dismutase (SOD1) by editing K63-linked polyubiquitin chains. J. Biol. Chem. 287, 28576–28585 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Chatterjee, A. et al. The role of the mammalian DNA end-processing enzyme polynucleotide kinase 3′-phosphatase in spinocerebellar ataxia type 3 pathogenesis. PLoS Genet. 11, e1004749 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  81. Gao, R. et al. Inactivation of PNKP by mutant ATXN3 triggers apoptosis by activating the DNA damage-response pathway in SCA3. PLoS Genet. 11, e1004834 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  82. Pfeiffer, A. et al. Ataxin-3 consolidates the MDC1-dependent DNA double-strand break response by counteracting the SUMO-targeted ubiquitin ligase RNF4. EMBO J. 36, 1066–1083 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Ashkenazi, A. et al. Polyglutamine tracts regulate beclin 1-dependent autophagy. Nature 545, 108–111 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Warrick, J. M. et al. Ataxin-3 suppresses polyglutamine neurodegeneration in Drosophila by a ubiquitin-associated mechanism. Mol. Cell. 18, 37–48 (2005).

    Article  CAS  PubMed  Google Scholar 

  85. Zeng, L. et al. The de-ubiquitinating enzyme ataxin-3 does not modulate disease progression in a knock-in mouse model of Huntington disease. J. Huntingtons Dis. 2, 201–215 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  86. Bichelmeier, U. et al. Nuclear localization of ataxin-3 is required for the manifestation of symptoms in SCA3: in vivo evidence. J. Neurosci. 27, 7418–7428 (2007). An important study demonstrating that expanded ATXN3 concentrates in the nucleus, where it may be most toxic to neurons.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Nóbrega, C. et al. RNA interference mitigates motor and neuropathological deficits in a cerebellar mouse model of Machado–Joseph disease. PLoS ONE. 9, e100086 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  88. Costa Mdo, C. et al. Toward RNAi therapy for the polyglutamine disease Machado–Joseph disease. Mol. Ther. 21, 1898–1908 (2013).

    Article  PubMed  CAS  Google Scholar 

  89. Aiba, Y. et al. Allele-selective inhibition of expression of huntingtin and ataxin-3 by RNA duplexes containing unlocked nucleic acid substitutions. Biochemistry 52, 9329–9338 (2013).

    Article  CAS  PubMed  Google Scholar 

  90. Moore, L. R. et al. Evaluation of antisense oligonucleotides targeting ATXN3 in SCA3 mouse models. Mol. Ther. Nucleic Acids 7, 200–210 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Teixeira-Castro, A. et al. Serotonergic signalling suppresses ataxin 3 aggregation and neurotoxicity in animal models of Machado–Joseph disease. Brain 138, 3221–3237 (2015). In this study, an unbiased screen suggested that antidepressants in the serotonin reuptake inhibitor class could be protective against the SCA3 disease protein.

    Article  PubMed  PubMed Central  Google Scholar 

  92. Costa, M. D. et al. Unbiased screen identifies aripiprazole as a modulator of abundance of the polyglutamine disease protein, ataxin-3. Brain 139, 2891–2908 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  93. Pavel, M. et al. CCT complex restricts neuropathogenic protein aggregation via autophagy. Nat. Commun. 7, 13821 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Cushman-Nick, M., Bonini, N. M. & Shorter, J. Hsp104 suppresses polyglutamine-induced degeneration post onset in a drosophila MJD/SCA3 model. PLoS Genet. 9, e1003781 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Simões, A. T. et al. Calpain inhibition reduces ataxin-3 cleavage alleviating neuropathology and motor impairments in mouse models of Machado–Joseph disease. Hum. Mol. Genet. 23, 4932–4944 (2014).

    Article  PubMed  CAS  Google Scholar 

  96. Liman, J. et al. CDK5 protects from caspase-induced Ataxin-3 cleavage and neurodegeneration. J. Neurochem. 129, 1013–1023 (2014).

    Article  CAS  PubMed  Google Scholar 

  97. Matos, C. A. et al. Ataxin-3 phosphorylation decreases neuronal defects in spinocerebellar ataxia type 3 models. J. Cell Biol. 212, 465–480 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Riess, O. et al. SCA6 is caused by moderate CAG expansion in the α1A-voltage-dependent calcium channel gene. Hum. Mol. Genet. 6, 1289–1293 (1997).

    Article  CAS  PubMed  Google Scholar 

  99. Zhuchenko, O. et al. Autosomal dominant cerebellar ataxia (SCA6) associated with small polyglutamine expansions in the α1A-voltage-dependent calcium channel. Nat. Genet. 15, 62–69 (1997).

    Article  CAS  PubMed  Google Scholar 

  100. Pietrobon, D. Calcium channels and channelopathies of the central nervous system. Mol. Neurobiol. 25, 31–50 (2002).

    Article  CAS  PubMed  Google Scholar 

  101. Kordasiewicz, H. B. et al. C-termini of P/Q-type Ca2+ channel α1A subunits translocate to nuclei and promote polyglutamine-mediated toxicity. Hum. Mol. Genet. 15, 1587–1599 (2006).

    Article  CAS  PubMed  Google Scholar 

  102. Du, X. et al. Second cistron in CACNA1A gene encodes a transcription factor mediating cerebellar development and SCA6. Cell 154, 118–133 (2013). This study reshaped thoughts about SCA6 pathogenesis with the discovery that the SCA6 disease gene is complex and encodes both the α1A subunit of the Cav2.1 calcium channel and a much smaller polyQ-containing transcription factor.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Unno, T. et al. Development of Purkinje cell degeneration in a knockin mouse model reveals lysosomal involvement in the pathogenesis of SCA6. Proc. Natl Acad. Sci. USA 109, 17693–17698 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  104. Yang, S., Li, X. J. & Li, S. Molecular mechanisms underlying spinocerebellar ataxia 17 (SCA17) pathogenesis. Rare Dis. 4, e1223580 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  105. Tsuji, S. Dentatorubral-pallidoluysian atrophy. Handb. Clin. Neurol. 103, 587–594 (2012).

    Article  PubMed  Google Scholar 

  106. Helmlinger, D. et al. Both normal and polyglutamine-expanded ataxin-7 are components of TFTC-type GCN5 histone acetyltransferase-containing complexes. Biochem. Soc. Symp. 73, 155–163 (2006).

    Article  CAS  Google Scholar 

  107. Lan, X. et al. Poly(Q) expansions in ATXN7 affect solubility but not activity of the SAGA deubiquitinating module. Mol. Cell. Biol. 35, 1777–1787 (2015). An important study showing that although expanded ATXN7 still functions in deubiquitination, its propensity to aggregate ultimately impairs DUB activity in cells.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. McCullough, S. D. & Grant, P. A. Histone acetylation, acetyltransferases, and ataxia — alteration of histone acetylation and chromatin dynamics is implicated in the pathogenesis of polyglutamine-expansion disorders. Adv. Protein Chem. Struct. Biol. 79, 165–203 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  109. Palhan, V. B. et al. Polyglutamine-expanded ataxin-7 inhibits STAGA histone acetyltransferase activity to produce retinal degeneration. Proc. Natl Acad. Sci. USA 102, 8472–8477 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  110. Furrer, S. A. et al. Spinocerebellar ataxia type 7 cerebellar disease requires the coordinated action of mutant ataxin-7 in neurons and glia, and displays non-cell-autonomous Bergmann glia degeneration. J. Neurosci. 31, 16269–16278 (2011). One of the more compelling studies that illustrates the complex interrelationship between neurons and non-neuronal cells in the precipitation of polyQ-mediated cerebellar degeneration.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  111. Barnes, J. A. et al. Abnormalities in the climbing fiber–Purkinje cell circuitry contribute to neuronal dysfunction in ATXN1[82Q] mice. J. Neurosci. 31, 12778–12789 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  112. Power, E. M., Morales, A. & Empson, R. M. Prolonged type 1 metabotropic glutamate receptor dependent synaptic signaling contributes to spino-cerebellar ataxia type 1. J. Neurosci. 36, 4910–4916 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  113. De Zeeuw, C. I. et al. Spatiotemporal firing patterns in the cerebellum. Nat. Rev. Neurosci. 12, 327–344 (2011).

    Article  CAS  PubMed  Google Scholar 

  114. Chopra, R. & Shakkottai, V. G. Translating cerebellar Purkinje neuron physiology to progress in dominantly inherited ataxia. Future Neurol. 9, 187–196 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Hourez, R. et al. Aminopyridines correct early dysfunction and delay neurodegeneration in a mouse model of spinocerebellar ataxia type 1. J. Neurosci. 31, 11795–11807 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  116. Dell'Orco, J. M. et al. Neuronal atrophy early in degenerative ataxia is a compensatory mechanism to regulate membrane excitability. J. Neurosci. 35, 11292–11307 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Kasumu, A. W. et al. Selective positive modulator of calcium-activated potassium channels exerts beneficial effects in a mouse model of spinocerebellar ataxia type 2. Chem. Biol. 19, 1340–1353 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  118. Hansen, S. T. et al. Changes in Purkinje cell firing and gene expression precede behavioral pathology in a mouse model of SCA2. Hum. Mol. Genet. 22, 271–283 (2013).

    Article  CAS  PubMed  Google Scholar 

  119. Egorova, P. A. et al. In vivo analysis of cerebellar Purkinje cell activity in SCA2 transgenic mouse model. J. Neurophysiol. 115, 2840–2851 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  120. Shakkottai, V. G. et al. Early changes in cerebellar physiology accompany motor dysfunction in the polyglutamine disease spinocerebellar ataxia type 3. J. Neurosci. 31, 13002–13014 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. Jayabal, S. et al. 4-Aminopyridine reverses ataxia and cerebellar firing deficiency in a mouse model of spinocerebellar ataxia type 6. Sci. Rep. 6, 29489 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  122. Keiser, M. S., Kordasiewicz, H. B. & McBride, J. L. Gene suppression strategies for dominantly inherited neurodegenerative diseases: lessons from Huntington's disease and spinocerebellar ataxia. Hum. Mol. Genet. 25, R53–R64 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  123. Miyazaki, Y. et al. An miRNA-mediated therapy for SCA6 blocks IRES-driven translation of the CACNA1A second cistron. Sci. Transl Med. 8, 347ra94 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  124. Scoles, D. R. et al. Antisense oligonucleotide therapy for spinocerebellar ataxia type 2. Nature 544, 362–366 (2017). One of several recent reports showing the promise of ASO-based disease gene silencing as a treatment for the polyQ SCAs. In this reference, the target is ATXN2, which the next reference shows is a therapeutic target in a TDP43 mouse model of ALS, providing further support for the concept that ATXN2 is an ALS risk factor.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  125. Becker, L. A. et al. Therapeutic reduction of ataxin 2 extends lifespan and reduces pathology in TDP-43 mice. Nature 544, 367–371 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. Chriboga, C. A. et al. Results from a Phase 1 study of ISIS-SMNRX in children spinal muscular atrophy. Neurology 12, 435–442 (2013).

    Google Scholar 

  127. Finkel, R. S. et al. Treatment of infantile-onset spinal muscular atrophy with nusinersen: a phase 2, open-label, dose-escalation study. Lancet 388, 3017–3026 (2016).

    Article  CAS  PubMed  Google Scholar 

  128. Miller, T. M. et al. An antisense oligonucleotide against SOD1 delivered intrathecally for patients with SOD1 familial amytrophic lateral sclerosis: a phase 1, randomized, first-in man study. Lancet Neurol. 12, 435–442 (2016).

    Article  CAS  Google Scholar 

  129. Jacobi, H. et al. Long-term disease progression in spinocerebellar ataxia types 1, 2, 3, and 6: a longitudinal cohort study. Lancet Neurol. 14, 1101–1108 (2015).

    Article  PubMed  Google Scholar 

  130. Reetz, K. et al. Genotype-specific patterns of atrophy progression are more sensitive than clinical decline in SCA1, SCA3, and SCA6. Brain 136, 905–917 (2013).

    Article  PubMed  Google Scholar 

  131. Adanyeguh, I. M. et al. In vivo neurometabolic profiling in patients with spinocerebellar ataxia types 1, 2, 3, and 7. Mov. Disord. 30, 662–670 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  132. Velázquez-Pérez, L. et al. Abnormal corticospinal tract function and motor cortex excitability in non-ataxic SCA2 mutation carriers: a TMS study. Clin. Neurophysiol. 127, 2713–2719 (2016).

    Article  PubMed  Google Scholar 

  133. Velázquez-Pérez, L. et al. Corticomuscular coherence: a novel tool to assess the pyramidal tract dysfunction in spinocerebellar ataxia type 2. Cerebellum 16, 602–606 (2017).

    Article  PubMed  CAS  Google Scholar 

  134. Southwell, A. L. et al. Ultrasensitive measurement of huntingtin protein in cerebrospinal fluid demonstrates increase with Huntington disease stage and decrease following brain huntingtin suppression. Sci. Rep. 5, 12166 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  135. Dansithong, W. et al. Ataxin-2 regulates RGS8 translation in a new BAC-SCA2 transgenic mouse model. PLoS Genet. 11, e1005182 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  136. Zu, T. et al. Recovery from polyglutamine-induced neurodegeneration in conditional SCA1 transgenic mice. J. Neurosci. 24, 8853–8861 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  137. Ikeda, Y. et al. Spectrin mutations cause spinocerebellar ataxia type 5. Nat. Genet. 38, 184–190 (2006).

    Article  CAS  PubMed  Google Scholar 

  138. Houlden, H. et al. Mutations in TTBK2, encoding a kinase implicated in tau phosphorylation, segregate with spinocerebellar ataxia type 11. Nat. Genet. 39, 1434–1436 (2007).

    Article  CAS  PubMed  Google Scholar 

  139. Waters, M. F. & Pulst, S. M. SCA13. Cerebellum 7, 165–169 (2008).

    Article  CAS  PubMed  Google Scholar 

  140. Chen, D. H. et al. Missense mutations in the regulatory domain of PKCγ: a new mechanism for dominant nonepisodic cerebellar ataxia. Am. J. Hum. Genet. 72, 839–849 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  141. Storey, E. & Gardner, R. J. Spinocerebellar ataxia type 15. Handb. Clin. Neurol. 103, 561–565 (2012).

    Article  PubMed  Google Scholar 

  142. Iwaki, A. et al. Heterozygous deletion of ITPR1, but not SUMF1, in spinocerebellar ataxia type 16. J. Med. Genet. 45, 32–35 (2008).

    Article  CAS  PubMed  Google Scholar 

  143. van de Leemput, J. et al. Deletion at ITPR1 underlies ataxia in mice and spinocerebellar ataxia 15 in humans. PLoS Genet. 3, e108 (2007).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  144. Verbeek, D. S. et al. Mapping of the SCA23 locus involved in autosomal dominant cerebellar ataxia to chromosome region 20p13-12.3. Brain 127, 2551–2557 (2004).

    Article  CAS  PubMed  Google Scholar 

  145. Hekman, K. E. et al. A conserved eEF2 coding variant in SCA26 leads to loss of translational fidelity and increased susceptibility to proteostatic insult. Hum. Mol. Genet. 21, 5472–5483 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Brusse, E. et al. Spinocerebellar ataxia associated with a mutation in the fibroblast growth factor 14 gene (SCA27): a new phenotype. Mov. Disord. 21, 396–401 (2006).

    Article  PubMed  Google Scholar 

  147. Mariotti, C. et al. Spinocerebellar ataxia type 28. Handb. Clin. Neurol. 103, 575–579 (2012).

    Article  PubMed  Google Scholar 

  148. Turcotte-Gauthier, M. et al. Expanding the clinical phenotype associated with ELOVL4 mutation: study of a large French–Canadian family with autosomal dominant spinocerebellar ataxia and erythrokeratodermia. JAMA Neurol. 71, 470–475 (2014).

    Article  PubMed  Google Scholar 

  149. Guo, Y.-C. et al. Spinocerebellar ataxia 35: novel mutations in TGM6 with clinical and genetic characterization. Neurology 83, 1554–1561 (2014).

    Article  CAS  PubMed  Google Scholar 

  150. Di Gregorio, E. et al. ELOVL5 mutations cause spinocerebellar ataxia 38. Am. J. Hum. Genet. 95, 209–217 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  151. Tsoi, H. et al. A novel missense mutation in CCDC88C activates the JNK pathway and causes a dominant form of spinocerebellar ataxia. J. Med. Genet. 51, 590–595 (2014).

    Article  CAS  PubMed  Google Scholar 

  152. Fogel, B. L., Hanson, S. M. & Becker, E. B. E. Do mutations in the murine ataxia gene TRPC3 cause cerebellar ataxia in humans? Mov. Disord. 30, 284–286 (2015).

    Article  CAS  PubMed  Google Scholar 

  153. Coutelier, M. et al. A recurrent mutation in CACNA1G alters Cav3.1 T-type calcium-channel conduction and causes autosomal-dominant cerebellar ataxia. Am. J. Hum. Genet. 97, 726–737 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  154. Depondt, C. et al. MME mutation in dominant spinocerebellar ataxia with neuropathy (SCA43). Neurol. Genet. 2, e94 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  155. White, M. et al. Transgenic mice with SCA10 pentanucleotide repeats show motor phenotypes and susceptibility to seizure: a toxic RNA gain-of-function model. J. Neurosci. Res. 90, 706–714 (2012).

    Article  CAS  PubMed  Google Scholar 

  156. Zu, T. et al. Non-ATG-initiated translation directed by microsatellite expansions. Proc. Natl Acad. Sci. USA 108, 260–265 (2011).

    Article  CAS  PubMed  Google Scholar 

  157. Cleary, J. D. & Ranum, L. P. Repeat associated non-ATG (RAN) translation: new starts in microsatellite expansion disorders. Curr. Opin. Genet. Dev. 26, 6–15 (2014). A comprehensive review of emerging evidence supporting the idea that RAN translation-generated peptides play a part in repeat-expansion-associated neurological disease.

    Article  CAS  PubMed  Google Scholar 

  158. Bañez-Coronel, M. et al. RAN translation in Huntington disease. Neuron 88, 667–677 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  159. Burright, E. N. et al. SCA1 transgenic mice: a model for neurodegeneration caused by an expanded CAG trinucleotide repeat. Cell 82, 937–948 (1995).

    Article  CAS  PubMed  Google Scholar 

  160. Huynh, D. P. et al. Nuclear localization or inclusion body formation of ataxin-2 are not necessary for SCA2 pathogenesis in mouse or human. Nat. Genet. 26, 44–50 (2000).

    Article  CAS  PubMed  Google Scholar 

  161. Lorenzetti, D. et al. Repeat instability and motor incoordination in mice with a targeted expanded CAG repeat in the Sca1 locus. Hum. Mol. Genet. 9, 779–785 (2000).

    Article  CAS  PubMed  Google Scholar 

  162. Watase, K. et al. A long CAG repeat in the mouse Sca1 locus replicates SCA1 features and reveals the impact of protein solubility on selective neurodegeneration. Neuron 34, 905–919 (2002).

    Article  CAS  PubMed  Google Scholar 

  163. Boy, J. et al. Reversibility of symptoms in a conditional mouse model of spinocerebellar ataxia type 3. Hum. Mol. Genet. 18, 4282–4295 (2009).

    Article  CAS  PubMed  Google Scholar 

  164. Dell'Orco, J. M., Pulst, S. M. & Shakkottai, V. G. Potassium channel dysfunction underlies Purkinje neuron spiking abnormalities in spinocerebellar ataxia type 2. Hum. Mol. Genet. https://doi.org/10.1093/hmg/ddx281 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

Download references

Acknowledgements

The authors thank J. Friedrich for designing the artwork that formed the foundation for the figures in this Review. Financial support for the authors' research was provided by the US National Institute of Neurological Disorders and Stroke, the US National Institutes of Health and the US National Ataxia Foundation.

Author information

Authors and Affiliations

Authors

Contributions

H.L.P., V.G.S., H.B.C. and H.T.O. researched data for the article, made substantial contributions to discussions of the content, wrote the article and reviewed and/or edited the manuscript before submission.

Corresponding authors

Correspondence to Henry L. Paulson or Harry T. Orr.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Related links

PowerPoint slides

Glossary

Pyramidal symptoms

Spasticity, increased reflexes and weakness caused by pathology in cerebral cortical primary motor neurons or their axonal projections.

Extrapyramidal symptoms

Rigidity, tremor, difficulty initiating movement and gait disturbances typical of parkinsonism or, occasionally, increased involuntary movement, such as chorea. These symptoms are secondary to pathology in pathways involving the basal ganglia and substantia nigra.

Ophthalmoplegia

Inability to move the eyes, secondary to pathology in the oculomotor cranial nerve nuclei (cranial nerves III, IV and VI) or the nerves themselves. In some cases, the pathology may be above the cranial nerve nuclei (supranuclear), in the voluntary gaze centres.

Oligonucleotide/oligosaccharide-binding fold

(OB fold). A structural feature found in many oligonucleotide-binding proteins.

Stress granules

Subcellular organelles that are major sites for the regulation of mRNA translation. These are non-membrane-bound aggregates composed of proteins and mRNA molecules that form in a reversible manner during cellular stress.

Processing bodies

(P-bodies). Cytoplasmic domains that contain collections of proteins involved in diverse processes, such as mRNA degradation, nonsense-mediated mRNA decay, translational repression and inhibitory-RNA-mediated gene silencing.

Cycloheximide

An antibiotic that interferes with protein synthesis by blocking translational elongation. The drug is used experimentally as a rapidly reversible means to block protein synthesis in vitro.

Heterogeneous nuclear ribonucleoprotein

Protein that forms nuclear complexes with RNA during gene transcription and post-translational modification of pre-mRNA.

Aggresome

A cytoplasmic collection of misfolded proteins that forms when the normal protein degradation system is overtaxed. This possibly protective mechanism may result in formation of the intracellular inclusions observed in neurodegenerative diseases.

Bicistronic mRNA

mRNA that has two open-reading frames, both of which are translated into a polypeptide. These polypeptides often are functionally related and may be regulated together.

Internal ribosomal entry site

(IRES). RNA structures that allow initiation of translation independent of the usual 5′-cap mechanism, thus enabling translation of protein in the middle of an mRNA.

SPT–ADA–GCN5 acetyltransferase complex

(SAGA complex). A histone acetyltransferase complex that regulates gene transcription by altering chromatin structure via histone modification. The complex is also involved in the biogenesis and nuclear export of mRNA.

Bergmann glia

A type of astrocyte found only in the cerebellar cortex. These cells have intimate contacts with Purkinje cell dendrites in the superficial layer of the cortex and are involved in the re-uptake of glutamate.

Negative allosteric modulator

A molecule that decreases the function of a receptor or an enzyme by binding to a locus that is not the active enzymatic site, typically by effecting a conformational change.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Paulson, H., Shakkottai, V., Clark, H. et al. Polyglutamine spinocerebellar ataxias — from genes to potential treatments. Nat Rev Neurosci 18, 613–626 (2017). https://doi.org/10.1038/nrn.2017.92

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrn.2017.92

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing